force&inducedreversalof&eliminations:+stressed

12
ForceInduced Reversal of Eliminations: Stressed Disulfide Bonds in Alkaline Solution Przemyslaw Dopieralski,* Jordi RibasArino,* Padmesh Anjukandi, Martin Krupicka, and Dominik Marx* Dr. P. Dopieralski, Dr. J. RibasArino, Dr. P. Anjukandi, Dr. M. Krupicka, Prof. Dr. D. Marx Lehrstuhl für Theoretische Chemie, RuhrUniversität Bochum 44780 Bochum (Germany) Email: [email protected] [email protected] [email protected] Dr. P. Dopieralski Permanent address: Faculty of Chemistry, University of Wroclaw JoliotCurie 14, 50383 Wroclaw (Poland) Dr. J. RibasArino Permanent address: Departament de Química Física and IQTCUB Universitat de Barcelona Av. Diagonal 645, 08028 Barcelona (Spain) Dr. M. Krupicka Current address: MaxPlanckInstitut für Chemische Energiekonversion Stiftstrasse 3436, 45470 Mülheim an der Ruhr (Germany)

Upload: others

Post on 20-Nov-2021

3 views

Category:

Documents


0 download

TRANSCRIPT

Force-­‐Induced  Reversal  of  -­‐Eliminations:  Stressed  Disulfide  Bonds  in  Alkaline  Solution  

   Przemyslaw  Dopieralski,*  Jordi  Ribas-­‐Arino,*  Padmesh  Anjukandi,  Martin  Krupicka,  and  Dominik  Marx*          Dr.  P.  Dopieralski,  Dr.  J.  Ribas-­‐Arino,  Dr.  P.  Anjukandi,  Dr.  M.  Krupicka,  Prof.  Dr.  D.  Marx  Lehrstuhl  für  Theoretische  Chemie,  Ruhr-­‐Universität  Bochum  44780  Bochum  (Germany)  E-­‐mail:    [email protected]    [email protected]  [email protected]    Dr.  P.  Dopieralski  Permanent  address:  Faculty  of  Chemistry,  University  of  Wroclaw  Joliot-­‐Curie  14,  50-­‐383  Wroclaw  (Poland)  Dr.  J.  Ribas-­‐Arino  Permanent  address:  Departament  de  Química  Física  and  IQTCUB  Universitat  de  Barcelona  Av.  Diagonal  645,  08028  Barcelona  (Spain)  Dr.  M.  Krupicka  Current  address:  Max-­‐Planck-­‐Institut  für  Chemische  Energiekonversion  Stiftstrasse  34–36,  45470  Mülheim  an  der  Ruhr  (Germany)                

  2  

Abstract    

Understanding  the   impact  of   tensile   forces  on  disulfide  bond  cleavage   is  not  only  

crucial   to   the   breaking   of   cross-­‐linkers   in   vulcanized  materials   such   as   strained  

rubber,  but  also  to  the  regulation  of  protein  activity  by  disulfide  switches.  By  using  

ab  initio   simulations   in   the   condensed   phase,   we   investigated   the   response   of  

disulfide  cleavage  by  β-­‐elimination  to  mechanical  stress.  We  reveal   that   the  rate-­‐

determining   first   step   of   the   thermal   reaction,  which   is   the   abstraction  of   the  β-­‐

proton,   is   insensitive   to   external   forces.   However,   forces   larger   than   about   1  nN  

were   found   to   reshape   the   free-­‐energy   landscape  of   the   reaction   so  dramatically  

that  a  second  channel  is  created,  where  the  order  of  the  reaction  steps  is  reversed,  

turning  β-­‐deprotonation  into  a  barrier-­‐free  follow-­‐up  process  to  C−S  cleavage.  This  

transforms  a  slow  and  force-­‐independent  process  with  second-­‐order  kinetics  into  

a  unimolecular  reaction  that  is  greatly  accelerated  by  mechanical  forces.  

 

     

  3  

The  impact  of  tensile  forces  on  the  kinetics  and  reaction  mechanisms  of  disulfide  

bond   reductions   has   been   intensely   investigated   in   recent   years.1-­‐15  Much   of   the  

current   interest   in   biochemistry   stems   from   the   increasingly   acknowledged   key  

role  played  by  the  reduction  of  so-­‐called  “disulfide  switches”,  commonly  found  

in  protein  regions  of  pronounced  mechanical  strain,  in  the  regulation  of  the  activity  

of  many  proteins.14-­‐18  Orthogonal  to  this  function  in  living  organisms  is  the  force‐

induced   breaking   of   sulfur‐based   cross‐linkers   in   vulcanized   rubber,  which   is  

essential   not   only   in   waste   rubber   recycling   (an   urgent   economic   and  

environmental   task   faced   by   industry   worldwide)19-­‐21   but   also   in   defining   the  

mechanical  fatigue  properties  of  rubber.22,  23  

 

In   landmark   single‐molecule   experiments   in   the   context   of   the   covalent  

mechanochemistry24-­‐28   of   disulfide   bond   breaking,   Fernandez   and   co-­‐workers  

probed   the  cleavage  of   such  bonds   in  aqueous  alkaline  solutions  as  a   function  of  

applied   tensile   force.3   These   force-­‐clamp   atomic   force   microscopy   (AFM)  

experiments  revealed  a  “mechanochemical  switch”  at  a  force  of  about  0.5  nN  above  

which  the  accelerating  effect  of  tensile  force  on  the  reaction  rate  was  considerably  

diminished.3   In   a   recent   computational   study,12   we   showed   that   the   reaction  

probed   in   these   experiments   is   S−S   bond   cleavage   triggered   by   a   nucleophilic  

attack   of   an   OH−   ion   onto   one   of   the   sulfur   atoms   by   an   SN2  mechanism.   Upon  

studying  the  response  of  this  reaction  pathway  to  tensile  stress,  we  disclosed  that  

the   mechanochemical   switch   originates   from   a   force‐induced   conformational  

rearrangement   of   the   disulfide   bridge,   which   drives   the   disulfide   into   a  

conformation   that   is   shielded   against   nucleophilic   attack.   Our   subsequent   work  

compared   the   force   response   of   the   SN2   mechanism,   which   is   preferred   at   low  

mechanical  stress,  to  other  scenarios  becoming  relevant  in  the  high-­‐force  regime.29  

 

When   it   comes   to   disulfide   bond   cleavage   reactions,   it   is  well   known   that   under  

certain   circumstances,30-­‐32   disulfides   can   undergo   β -­‐elimination   in   alkaline  

solution.   This   requires   investigations   into   the   response   of   this   more   complex  

reaction  pathway  to  mechanical  stress  to  achieve  a  detailed  understanding  of  the  

mechanochemistry   of   disulfide   bridges.   Herein,   based   on   ab  initio   molecular  

  4  

dynamics   studies   in   the   condensed   phase   enhanced   with   metadynamics  

sampling,33  we  explore  how  the  free‐energy  landscape  associated  with  the  β‐

elimination   channel   gets   transformed  under   the   action  of   external   tensile   stress.  

Unexpectedly,  we  disclose  significant   force-­‐induced  changes   in   the  mechanism  at  

intermediate   forces,  which   reveals   a   novel   reaction   channel   aside   from   the   slow  

and   force-­‐insensitive   bimolecular   process   in   terms   of   a   greatly   accelerated  

unimolecular  process.  

 

The  model  system  introduced  here  to  explore  the  mechanical  response  of  the  β‐

elimination   scenario   greatly   exceeds   our   previously   studied   minimal   diethyl  

disulfide  model  in  complexity12,  29  (Figure  1,  inset),  and  the  model  compound  was  

hence   solvated   in   a   spacious   bulk   water   periodic   supercell   that   also   hosts   the  

attacking   fully   solvated   OH−   (aq).   In   particular,   the   disulfide   model   employed  

herein  exhibits  one  of  its  sulfur  atoms  linked  to  a  large  organic  moiety  in  which  the  

β‐carbon   atom   of   the   disulfide   is   bonded   to   two   fragments   excised   from   the  

protein  backbone  used   in   the  AFM  experiments.3  This  ensures   that   the  acidity  of  

the   β-­‐proton   (see   Figure  1,   inset)   is   properly   represented,   whereas   the   far‐

distant   sulfur   site   is   saturated   using   a   methyl   group.   Two   so-­‐called   collective  

variables   (CVs)   were   introduced   to   span   the   free‐ energy   landscape   that  

characterizes  β‐elimination  (see  the  Supporting  Information  for  details):  The  first  

one  is  the  coordination  number,  CNC−S,  between  the  carbon  and  sulfur  atoms  of  the  

C−S  bond  (denoted  by  CV1  in  Figure  2).  CV2  is  the  difference  in  the  coordination  

number  between  the  β-­‐hydrogen  and  the  carbon  atom  to  which  it  is  linked  in  the  

disulfide  molecule,  CNC−Hβ,   and   the  coordination  number  of   the  β-­‐hydrogen  atom  

with  respect  to  the  oxygen  atom  of  the  hydroxide  ion,  that  is,  CNC−Hβ  –  CNC−Hβ  -­‐  OH-­‐  

  5  

 Figure  1.  Force  dependence  of  the  activation  free  energy  for  β-­‐elimination  of  the  disulfide  bond   (see   inset),   corresponding   to   the   rate-­‐determining   bimolecular   and   unimolecular  first   step   of   the   reaction   according   to   the   carbanion   (•)   and   carbocation   (▪)   pathways,  which  are  preferred  at  low  and  high  force,  respectively.  The  inset  shows  the  large  disulfide  model  currently  employed  together  with  the  attacking  hydroxide  ion  (dashed  red  arrow);  all   water   molecules   were   removed   for   clarity.   The   β-­‐hydrogen   atom   (green)   gets  abstracted  as  a  proton  by  OH−  (aq).  C  black,  H  gray,  N  blue,  O  red,  S  yellow.  Arrows  indicate  the  C  atoms  on  which  constant  mechanical   forces  of  magnitude  F  are  exerted  collinearly  along  a  fixed  direction  in  space.      

The  free  energy  surface  (FES)  obtained  for  β-­‐elimination  in  the  thermal  limit  (i.e.,  

in   the  absence  of  any  external  mechanical   force)  demonstrates   that   this   reaction  

follows  a  two‐step  mechanism  (see  Figure  2 A).  In  the  rate-­‐determining  first  step  

(which  connects  the  reactant  R  to  product  P1  via  transition  state  TS1),  the  OH−  ion  

removes   the   β-­‐proton   of   the   disulfide   by   overcoming   a   significant   free-­‐energy  

barrier  of  ΔA≠≈35  kcal mol−1  (see  Figure  1  at  F=0  nN).  The  resulting  product  P1  

is  a  short-­‐lived  carbanion  intermediate  where  the  attacking  OH−  species,  which  has  

been  turned   into  a  water  molecule  by  H  abstraction,  creates  an  H-­‐bonded  bridge  

that  connects  one  of  the  sulfur  atoms  with  an  oxygen  atom  of  the  amide  bond  (see  

Figure   2 A,   inset,   P1).   The   second   step   (P1→TS2→P2   in   Figure   2 A)   consists   of  

simple  C−S  bond  cleavage  with  a  vanishingly  small  and  force-­‐independent  barrier  

on   the   order   of   1   kcal mol−1;   the   thereby   generated   disulfide   anion   eventually  

accepts  a  proton  from  water.  Overall,   this   is  a  two‐step  “irreversible  carbanion”  

mechanism,34,  35   (E1cB)I,   that   is   characterized   by   second‐order   kinetics,   where  

  6  

the   first   bimolecular   step—involving   the   conjugate   base   and   yielding   carbanion  

P1—is  rate‐determining  and  essentially  irreversible  given  that  the  second  step  is  

very   fast.   Indeed,   this   finding   is   in   agreement  with   general   expectations   for   the  

purely  thermal  reaction  as  electron‐withdrawing  groups  in  the  β-­‐position,  such  

as   the   CONH2   group   in   our   case,   shift   β -­‐eliminations   towards   the   E1cB  

mechanism.35   Moreover,   strong   bases   and   high   base   concentrations   also   favor  

E1cB,35  which  is  certainly  the  case  for  our  aqueous  alkaline  solution.  

 Figure   2.  A)  Free-­‐energy   landscape   for  β-­‐elimination  (see  Figure  1)   in   the   thermal   limit  (i.e.,   at   F=0   nN)   where   only   the   carbanion   pathway   is   operating.   B)   Force-­‐transformed  

  7  

free-­‐energy  landscape  for  β-­‐elimination  (see  Figure  1)  at  a  constant  force  of  F=1  nN  where  the   carbocation   channel   is   energetically   preferred   over   the   still   operating   carbanion  pathway.   Red   and   blue   arrows   indicate   the   carbanion   and   carbocation   pathways,  respectively,   whereas   R,   Pn,   and   TSn   denote   reactants,   intermediates/products,   and  transition  states,  respectively,  for  which  representative  snapshots  sampled  from  ab  initio  metadynamics  are  shown.    

Having   successfully   deciphered   the   thermal  β-­‐elimination   reaction   mechanism,  

we   focused   on   the   effect   of   finite  mechanical   tensile   forces   on   this   process.   This  

investigation  was  performed  by  means  of  isotensional  metadynamics  sampling28,36  

carried  out  in  the  presence  of  different,  but  constant  external  forces  (see  Figure  1).  

For   each   fixed   value   of   the   external   force,   the   isotensional   metadynamics  

simulations   lead   to   a   different   so‐called   “force‐transformed   FES”,36   which  

provides  the  proper  reaction  landscape  including  the  mechanical  distortion.28  

 

At  a  force  of  0.5  nN,  the  entire  mechanism  remains  unchanged  when  compared  to  

the  purely   thermally  activated  process  (see   the  corresponding   force-­‐transformed  

FES  in  the  Supporting  Information,  Figure  S3).  Interestingly,  however,  the  barrier  

ΔA≠   for   the   rate-­‐determining   first   step   remains   roughly   constant   according   to  

Figure  1,  which  means  that  the  external  force  does  not  result  in  any  acceleration  of  

the   overall   reaction.   Larger   forces   do   not   lead   to   a   reduction   of   this   free-­‐energy  

barrier   either   (Figure   1).   Our   simulations   thus   clearly   reveal   that   the   rate-­‐

determining   step   of   the   (E1cB)I   process,   that   is,  β-­‐proton   abstraction,   is   fully  

decoupled  from  the  applied  mechanical  force.  Other  examples  of  force-­‐insensitive  

chemical  processes  whose  reaction  coordinates  are  orthogonal   to  the  mechanical  

coordinates  can  be  found  in  Refs.  5  and  37-­‐39.  

 

A  dramatic  change  of  scenario   is  observed  upon  reaching  an  intermediate  tensile  

force  of  1  nN.  Indeed,  Figure  2 B  discloses  that  the  force-­‐transformed  FES  enables  a  

different  reaction  channel:  First,  the  carbocation  P3  is  obtained  by  cleaving  the  C−S  

bond  without  involvement  of  the  base,  OH−(aq),  which  only  subsequently  abstracts  

the   β-­‐proton   (R→P3→P2);   the   disulfide   anion   generated   after   the   first   step  

eventually  stabilizes  itself  by  abstracting  a  proton  from  the  solvent.  The  activation  

free   energy   of   the   rate‐determining   first   step   at   F=1.0  nN,   ΔA≠≈28   kcal mol−1  

according   to   Figure   1,   is   much   smaller   than   that   associated   with   the   rate‐

  8  

determining   first   step   of   the   (E1cB)I   process,   which   is   37   kcal mol−1   for   β‐H  

abstraction   by   OH−   (aq)   at   1   nN.   However,   the   generation   of   carbocations   in  

aqueous   solution   appears   to   be   an   utmost   unrealistic   scenario.   Indeed,   our  

simulations  show  that  no  carbocations  are  generated  in  the  limit  of  the  thermally  

activated  reaction  (i.e.,  at  F=0  nN,  see  Figure  2 A),  and  that  their  generation  is  also  

not  feasible  at  moderate  forces  below  approximately  1  nN  (see  Figure  1).  So,  why  

is  carbocation  generation  enabled  by  sufficiently  large  mechanical  forces?  

 

Mechanistic  analysis  of  the  novel  scenario  shows  that  the  same  final  product  P2  as  

before   is   obtained,   yet   by   a   vastly   different  mechanism.   It  was   classified   to   be   a  

two-­‐step  E1  process35   as   the   rate-­‐determining   first   step   leads,   in   a  unimolecular  

process   (R→P3),   to  a   carbocation   that   readily   loses   the  β-­‐proton   (see  P3→P2   in  

Figure  2 B),  therefore  leading  to  first‐order  kinetics.  The  picture  underlying  this  

scenario   is   that   the   internal   strain   generated   by   mechanically   stretching   the  

molecule  is  so  large  that  the  covalent  bond  is  simply  ripped  apart  (R→P3)  before  

any   other   process   occurs.   The   nascent   carbocation   is   short‐lived   and   stabilizes  

itself   quickly   after   its   generation   by   detaching   the   β-­‐proton   into   the   basic  

solution.  This  view  is  supported  by  the  observation  that  the  rate‐determining  R→

P3  process,   and   thus   the  entire  E1  pathway,   is   very   responsive   to   tensile   forces:  

The  activation  free  energy,  ΔA≠,   is  reduced  to  half  from  about  28  kcal mol−1  at  1  

nN  to  approximately  14  kcal mol−1  at  2  nN.  Moreover,  Mulliken  population  analysis  

(see  Figure  S1)  reveals  that  the  α-­‐carbon  atom  involved  in  the  C−S  bond  cleavage  

of   P3   intermittently   acquires   a   positive   partial   charge.   Last   but   not   least,  

reference‐coupled   cluster   calculations   of   C−S   bond   breaking   using   a   reduced  

model  embedded  in  continuum  water  provide  no  support  for  a  homolytic  diradical  

mechanism   instead   of   the   heterolytic   ionic   pathway   involving   the   reported  

carbocationic   intermediate.   In   conclusion,   the   E1   reaction   channel   favored   at  

sufficiently   large   forces   not   only   implies   a   significant   acceleration   of   a   hitherto  

force‐insensitive  reaction,  but  also  a   force‐induced  change   from  a  bimolecular  

to  a  unimolecular  process  and  from  second-­‐order  to  first-­‐order  reaction  kinetics.  

 

  9  

Given  the  vastly  different  reaction  channel  that  is  preferred  at  forces  beyond  1  nN,  

our   final   analysis   addresses   the   nature   of   the   elementary   β-­‐proton   abstraction  

step.   Figure   3   nicely   reveals   that   C−S   cleavage   (dark‐brown   line)   occurs  

after/before   β-­‐H   abstraction   by   OH−   (green   lines)   in   the   carbanion/carbocation  

mechanism  depicted  in  panels  B/D  at  F=1.0  nN.  The  OH−  (aq)  hydration  shell  (red  

line)   is   more   flexible   in   the   carbanion   scenario,   whereas   mostly   two   water  

molecules   remain  hydrogen‐bonded  along   the  carbocation  pathway  (see  panels  

A/C,   respectively).   Moreover,   the   β-­‐proton-­‐abstracting   OH−   (aq)   forms   a   stable  

hydrogen  bond  to  one  of   the  NH  groups  of   the  reactant  molecule  (blue   line).  The  

fact  that  this  H-­‐bond  is  broken  shortly  after  the  β-­‐proton  is  abstracted,  but  largely  

locked   before,   suggests   that   this   hydrogen   bond   keeps   the   nucleophile   at   the  

reaction  center.  Indeed,  such  an  hydrogen‐bonded  intermediate  state  (HBIS)  has  

been   observed   for   base‐catalyzed   peptide   hydrolysis.38   Here,   only   a   slight  

migration  of  OH−  from  such  an  HBIS  configuration  to  the  C−Hβ  group  is  required  to  

achieve  the  elementary  elimination  step.  

 

 Figure   3.   Elementary   elimination   step   involving   b-­‐proton   abstraction   by   OH-­‐   (aq)   at   a  constant  force  of  F  =  1  nN  along  the  carbanion  and  carbocation  reaction  channels  (top  and  bottom,   respectively).   The   left   panels   depict   configuration   snapshots   of   the   reactive  complex  when  OH-­‐  (aq)  is  about  to  abstract  the  b-­‐proton  (green  dotted  line)  and  to  break  

  10  

its  stabilizing  hydrogen  bond  to  H-­‐N-­‐(red  dotted  line)  together  with  a  contour  plot  of  the  corresponding  spatial  distribution  functions  of  the  oxygen  site  of  OH  in  the  N-­‐H···Hb  plane  sampled  shortly  before,  during,  and  after  Hb  abstraction  and  thus  H2O  formation.  The  bulk  water  environment  present  in  the  simulations  is  not  shown.  The  panels  on  the  right  show  in  their  lower  parts  the  evolution  of  the  following  distances  along  the  two  pathways:  dark  brown:  C-­‐S  distance;  blue  and  green  lines:  distance  from  the  O  site  of  OH-­‐  to  the  H  sites  of  the  H-­‐N-­‐  and  Hβ-­‐C-­‐  groups,   respectively;   red   line:  number  of  water  molecules  hydrogen-­‐  bonded  to  OH-­‐  (aq).  The  top  parts  show  the  coordination  of  OH-­‐  (aq)  with  respect  to  the  H  sites  of  the  H-­‐N-­‐  and  Hβ  –C-­‐  groups  using  again  blue  and  green  lines,  where  the  final  jumps  to   zero   and   unity   correspond   to   the   termination   of   the  HBIS   (see   text)   and   to   β-­‐proton  abstraction,  respectively;  a  simple  distance  criterion  of  R<2.4  Å  was  used  to  visualize  the  hydration  and  coordination  numbers.  Note  that  OH-­‐  (aq)  changes  to  H2O  (aq)  along  the  red  line  shortly  before  the  green  and  blue  lines  cross.    

 

Acknowledgements

We are grateful to partial financial support from the Deutsche Forschungsgemeinschaft (Reinhart Koselleck Grant “Understanding Mechanochemistry”, MA 1547/9) and Cluster of Excellence “RESOLV”, EXC 1069)), the Alexander von Humboldt Stiftung (Humboldt Fellowship to J.R.‐A.), and the Spanish Government (Ramón y Cajal Fellowship to J.R.‐A.), as well as by the National Science Center Poland (2014/13/B/ST4/05009), and the Ministry of Science and Higher Education Poland (627/STYP/9/20l4) for fellowships to P.D. We also gratefully acknowledge the Gauss Centre for Supercomputing (GCS) for providing computing time for a GCS Large Scale Project on JUQUEEN at the Jülich Supercomputing Centre (JSC) as well as additional computational support by HPC‐RESOLV, BOVILAB@RUB, the Rechnerverbund NRW, the Wroclaw Supercomputer Center (WCSS), the Galera‐ACTION Cluster, and the Academic Computer Center in Gdańsk (CI TASK).

 

 

 

   

  11  

References  

 

[1] A. P. Wiita, S. R. K. Ainavarapu, H. H. Huang, J. M. Fernandez, Proc. Natl. Acad. Sci. USA 2006, 103, 7222 – 7227.

[2] A. P. Wiita, R. Perez-Jimenez, K. A. Walther, F. Gräter, B. J. Berne, A. Holmgren, J.M. Sanchez-Ruiz, J.M. Fernandez, Nature 2007, 450, 124 – 127.

[3] S. Garcia-Manyes, J. Liang, R. Szoszkiewicz, T. L. Kuo, J. M. Fernandez, Nat. Chem. 2009, 1, 236 – 242.

[4] J. Liang, J. M. Fernandez, ACS Nano 2009, 3, 1628 – 1645.���

[5] T. J. Kucharski, Z. Huang, Q. Z. Yang, Y. Tian, N. C. Rubin, C. D. Concepcion, R. Boulatov, Angew. Chem. Int. Ed. 2009, 48,

7040 – 7043 ; Angew. Chem. 2009, 121, 7174 – 7177.���

[6] W. Li, F. Gräter, J. Am. Chem. Soc. 2010, 132, 16790–16795.

[7] M.F.Iozzi, T.Helgaker, E.Uggerud, J.Phys.Chem.A2011,115,2308 – 2315.���

[8] S. Keten, C.-C. Chou, A. C. van Duin, M. J. Buehler, J. Mech. Behav. Biomed. Mater. 2012, 5, 32 – 40.

[9] F. Hofbauer, I. Frank, Chem. Eur. J. 2012, 18, 16332 – 16338.

[10] I. Popa, P. Kosuri, J. Alegre-Cebollada, S. Garcia-Manyes, J. M. Fernandez, Nat. Protoc. 2013, 8, 1261 – 1276.

[11] Y. Tian, T. J. Kucharski, Q. Z. Yang, R. Boulatov, Nat. Commun. 2013, 4, 2538.

[12] P. Dopieralski, J. Ribas-Arino, P. Anjukandi, M. Krupicka, J. Kiss, D. Marx, Nat. Chem. 2013, 5, 685 – 691.

[13] Y. Li, A. Nese, K. Matyjaszewski, S. S. Sheiko, Macromolecules, 2013, 46, 7196 – 7201.

[14]  P.  Anjukandi,  P.  Dopieralski,  J.  Ribas-­‐Arino,  D.  Marx,  PLoS  One  2014,  9,  

e108812.  

[15]  B.  Zhou,  I.  B.  Baldus,  W.  Li,  S.  A.  Edwards,  F.  Gräter,  Biophys.  J.  2014,  107,  672  

–  681.  

[16]  P.  J.  Hogg,  Trends  Biochem.  Sci.  2003,  28,  210  –  214.  

[17]  J.  Ahamed,  H.  H.  Versteeg,  M.  Kerver,  V.  M.  Chen,  B.  M.  Mueller,  P.  J.  Hogg,  W.  

Ruf,  Proc.  Natl.  Acad.  Sci.  USA  2006,  103,  13932–  13937.  

[18]  M.  A.  Wouters,  S.  W.  Fan,  N.  L.  Haworth,  Antioxid.  Redox  Signaling  2010,  12,  53  

–  91.  

  12  

[19]  X.  Guo,  D.  Xiang,  G.  Duan,  P.  Mou,  Waste  Manage.  2010,  30,  4  –10.  

[20]  X.  Zhang,  Z.  Lu,  D.  Tian,  H.  Li,  C.  Lu,  J.  Appl.  Polym.  Sci.  2013,  127,  4006  –  4014.  

[21]  H.  Rust,  US8957119  B2,  2015.  

[22]  W.  V.  Mars,  A.  Fatemi,  Rubber  Chem.  Technol.  2004,  77,  391  –412.  

[23]  B.  N.  J.  Persson,  O.  Albohr,  G.  Heinrich,  H.  Ueba,  J.  Phys.  Condens.  Matter  2005,  

17,  R1071  –  R1142.  

[24]  M.  K.  Beyer,  H.  Clausen-­‐Schaumann,  Chem.  Rev.  2005,  105,  2921  –  2948.  

[25]  M.  M.  Caruso,  D.  A.  Davis,  Q.  Shen,  S.  A.  Odom,  N.  R.  Sottos,  S.  R.  White,  J.  S.  

Moore,  Chem.  Rev.  2009,  109,  5755  –  5798.  

[26]  A.  L.  Black,  J.  M.  Lenhardt,  S.  L.  Craig,  J.  Mater.  Chem.  2011,  21,  1655–  1663.    [27]  Z.  Huang,  R.  Boulatov,  Chem.  Soc.  Rev.  2011,  40,  2359–  2384.    [28]  J.  Ribas-­‐Arino,  D.  Marx,  Chem.  Rev.  2012,  112,  5412–  5487.    [29]  P.  Dopieralski,  J.  Ribas-­‐Arino,  P.  Anjukandi,  M.  Krupicka,  D.  Marx,  unpublished  results.    [30]  M.  Friedman,  J.  Agric.  Food  Chem.  1999,  47,  1295–  1319.    [31]  S.  Cohen,  C.  Price,  J.  Vlasak,  J.  Am.  Chem.  Soc.  2007,  129,  6976–  6977.    [32]  B.  Lagrain,  K.  de  Vleeschouwer,  I.  Rombouts,  K.  Brijs,  M.  Hendrickx,  J.  Delcour,  J.  Agric.  Food  Chem.  2010,  58,  10761–  10767.    [33]  D.  Marx,  J.  Hutter,  Ab  Initio  Molecular  Dynamics:  Basic  Theory  and  Advanced  Methods,  Cambridge  University  Press,  Cambridge,  2009.    [34]  F.  G.  Bordwell,  Acc.  Chem.  Res.  1972,  5,  374–  381.    [35]  M.  B.  Smith,  March's  Advanced  Organic  Chemistry:  Reactions,  Mechanisms  and  Structure,  7th  ed.,  Wiley,  Hoboken,  2013.    [36]  P.  Dopieralski,  J.  Ribas-­‐Arino,  D.  Marx,  Angew.  Chem.  Int.  Ed.  2011,  50,  7105–  7108;    [37]  J.  Ribas-­‐Arino,  M.  Shiga,  D.  Marx,  Chem.  Eur.  J.  2009,  15,  13331–  13335.    [38]  F.  Xia,  A.  Baranowska,  S.  Cheng,  F.  Gräter,  J.  Phys.  Chem.  B  2011,  115,  10126–  10132.    [39]  R.  Groote,  B.  M.  Szyja,  F.  A.  Leibfarth,  C.  J.  Hawker,  N.  L.  Doltsinis,  R.  P.  Sijbesma,  Macromolecules  2014,  47,  1187–  1192.